Population Coding of Reward Magnitude in the Orbitofrontal Cortex of the Rat

Size: px
Start display at page:

Download "Population Coding of Reward Magnitude in the Orbitofrontal Cortex of the Rat"

Transcription

1 8590 The Journal of Neuroscience, August 20, (34): Behavioral/Systems/Cognitive Population Coding of Reward Magnitude in the Orbitofrontal Cortex of the Rat Esther van Duuren, Jan Lankelma, and Cyriel M. A. Pennartz Cognitive and Systems Neuroscience Group, Center for Neuroscience, Swammerdam Institute for Life Sciences, University of Amsterdam, 1098 SM Amsterdam, The Netherlands Although single-cell coding of reward-related information in the orbitofrontal cortex (OFC) has been characterized to some extent, much less is known about the coding properties of orbitofrontal ensembles. We examined population coding of reward magnitude by performing ensemble recordings in rat OFC while animals learned an olfactory discrimination task in which various reinforcers were associated with predictive odor stimuli. Ensemble activity was found to represent information about reward magnitude during several trial phases, namely when animals moved to the reward site, anticipated reward during an immobile period, and received it. During the anticipation phase, Bayesian and template-matching reconstruction algorithms decoded reward size correctly from the population activity significantly above chance level (highest value of 43 and 48%, respectively; chance level, 33.3%), whereas decoding performance for the reward delivery phase was 76 and 79%, respectively. In the anticipation phase, the decoding score was only weakly dependent on the size of the neuronal group participating in reconstruction, consistent with a redundant, distributed representation of reward information. In contrast, decoding was specific for temporal segments within the structure of a trial. Decoding performance steeply increased across the first few trials for every rewarded odor, an effect that could not be explained by a nonspecific drift in response strength across trials. Finally, when population responses to a negative reinforcer (quinine) were compared with sucrose reinforcement, coding in the delivery phase appeared to be related to reward quality, and thus was not based on ingested liquid volume. Key words: reward magnitude; neuronal ensembles; orbitofrontal cortex; tetrodes; Bayesian method; template matching Introduction The orbitofrontal cortex (OFC) is thought to contribute to the guidance of goal-directed behavior through the formation of neural representations of predicted outcomes. Studies examining single-unit activity in OFC during operant behavior demonstrated that firing rates are modulated by the motivational value of stimuli and represent reward-predictive information during anticipation of various types or amounts of reward (Thorpe et al., 1983; Lipton et al., 1999; Schoenbaum et al., 1999, 2003; Tremblay and Schultz, 1999; Hikosaka and Watanabe, 2000; Yonemori et al., 2000; Wallis and Miller, 2003; Roesch and Olson, 2004; Ichihara-Takeda and Funahashi, 2006; Padoa-Schioppa and Assad, 2006; Roesch et al., 2006; Ramus et al., 2007; Simmons and Richmond, 2008). These single-cell studies have not shown how predicted or actual rewards are dynamically represented by populations in OFC. This question is especially relevant for understanding how other, connected brain areas may read out population activity from this structure (Pouget et al., 2000; Wu and Amari, 2005). Until now, population coding analyses have been Received Dec. 15, 2007; revised July 3, 2008; accepted July 9, ThisworkwassupportedbyTheNetherlandsOrganizationforScientificResearch(NWO)Grant ,NWO Grant , and Besluit Subsidies Investeringen Kennisinfrastructuur (SenterNovem) Grant We thank David Redish and Peter Lipa for providing the cluster cutting software, Els Velzing for help with graphical illustrations, and our colleagues for their comments on this manuscript. Correspondence should be addressed to Cyriel M. A. Pennartz, Kruislaan 320, 1098 SM Amsterdam, The Netherlands. pennartz@science.uva.nl. DOI: /JNEUROSCI Copyright 2008 Society for Neuroscience /08/ $15.00/0 mainly used for reconstruction of arm movement direction during a reaching task (e.g., Georgopoulos et al., 1986), reconstruction of an animal s environmental location (Wilson and Mc- Naughton, 1993), and prediction of behavioral responses (Laubach et al., 2000; Baeg et al., 2003). Only a few studies have thus far described ensemble activity within OFC. Gutierrez et al. (2006) found that ensemble activity in rat OFC discriminated between sucrose and water reward, but this distinction was made in a free-licking situation, thus without predictive cues and a contingent operant response. In a sensory discrimination task in rats, Schoenbaum and Eichenbaum (1995) showed that, during stimulus sampling, OFC population activity represented expectation of a reward presented in the following trial. However, it remains unknown how actual and predicted rewards are represented in the population within a specific trial phase during an operant task, and whether such a population code would be specific for different trial phases. Here it is of special interest to compare representations when the animal anticipates reinforcement or when he receives it. For the reinforcer consumption phase, we also asked whether the observed ensemble coding is related to reward quality or can be explained by coding of the ingested volume of liquid reinforcer. In addition, we examined whether reward magnitude is represented in OFC in a sparse or redundant manner, i.e., by a few highly specifically tuned cells or in a broadly distributed way (cf. Narayanan et al., 2005). Finally, we addressed the dynamics by which consistent coding develops as the learning task progresses. It has been hardly feasible to ex-

2 van Duuren et al. Population Coding of Reward Magnitude J. Neurosci., August 20, (34): amine learning-related changes on a trial-by-trial basis in singleunit recordings from OFC, because single-unit firing patterns show great variability from trial to trial, making it difficult to track systematic changes during learning. To address these various aspects of population coding, ensemble recordings were made from rat OFC during a five-odor discrimination go/nogo task, in which odors were predictive for various amounts of an appetitive sucrose solution, for no reward or an aversive quinine solution. Materials and Methods Subjects All experiments were approved by the Animal Experimentation Committee of the Royal Netherlands Academy of Arts and Sciences and were performed in accordance with the National Guidelines for Animal Experimentation. Data were collected in 17 sessions recorded from six male Wistar rats [Harlan CPB; sessions partially overlapped with those presented in a single-unit study (van Duuren et al., 2007)]. Animals were weighed and handled daily and socially housed in standard type 4 macrolon cages under a reversed 12 h light/dark cycle (dimmed red light at 7:00 A.M.). Weight at the time of surgery was g. After surgery, animals were housed individually in a larger cage (1 1 1 m) under similar conditions. Food was available ad libitum, but animals were water restricted. They had access to water with a variable delay after the end of the recording session in the home cage for a 2 hperiod. Behavior Apparatus Recordings were made in a black Plexiglas operant chamber ( cm), placed in a sound-attenuated and electrically shielded box. The front panel contained on the right side a light signaling trial onset and an odor sampling port, and on the left side a delivery well for fluid reinforcement. Responses made by the animal in the sampling port and fluid well were registered by an infrared beam transmitter and detector. Behavioral events during task performance and data collection were controlled by a computer. Odor delivery was controlled by a system of solenoid valves and flow meters. To prevent mixing up of odors within the system, separate delivery lines for each odor were present. The different types of fluid reinforcement were delivered with separate fluid lines. Fluid delivery was gravity driven, with a tap and valves controlling the flow and amount of fluid delivered. The odorants (Tokos BV) were separated into different families, i.e., woody, fruity, herbal, citrus, and floral. Each set of odors used in a discrimination session contained one odor from each family. In addition, no single family of odors was preferentially associated with a particular trial outcome. Behavioral paradigm When animals were habituated to the recording chamber, the training on the behavioral procedure of the five-odor olfactory discrimination gono/go task started. In this task, odors signaled whether a go response resulted in a particular amount of a positive reinforcement or in a negative one. We used five different odors in each session, each of which was uniquely predictive throughout the session of either an amount of an appetitive sucrose solution (10% sucrose in water, i.e., 0.05, 0.15, and 0.30 ml), no reward (nonreinforced condition), or an aversive quinine solution (0.15 ml of M quinine in water). A particular odor was never used in more than one session. Criterion for behavioral performance was set at 15 trials per positively rewarded trial type, because it was difficult to reliably obtain more trials because of the larger reward volumes. When animals reached this criterion during training, they were implanted with a multitetrode array ( hyperdrive ), and recordings started. For each recording session, a new set of five odors was used. During the task, in which odors were presented pseudorandomly, the onset of the trial light indicated that animals could initiate a trial by making an odor poke. If no odor poke was made within 15 s after onset of the light, the light switched off and the intertrial interval started (with a variable duration of s). After initiation of an odor poke the trial light switched off after 500 ms, followed 500 ms later by the odor presentation. This interval was included to prevent the animal from moving during cue sampling. Odor sampling itself was required to last at least 1 s. After retraction from the odor sampling port or whenever a maximal duration for odor sampling (10 s) was exceeded, odor presentation was terminated. Premature retraction from the odor sampling port (odor poke shorter than the minimal duration of 2 s) resulted in the onset of the intertrial interval. The nose poke in the fluid well marked the onset of an immobile waiting period of 1.5 s, after which reinforcement was delivered. Whenever reinforcement was delivered, animals had 10 s to consume the reward, after which the intertrial interval started. Incorrect (go) responses after sampling an odor predictive of quinine or the nonrewarded contingency had no further programmed consequences. Furthermore, no other, separate reinforcer was applied during these trials. The behavioral sequence comprising the departure from the sampling port to the fluid well, including nose entry and waiting period in the well, will be referred to as the go response. Surgery, electrophysiology, and histology Animals were anesthetized with 0.08 ml/100 g of Hypnorm intramuscularly and 0.04 ml/100 g of Dormicum subcutaneously (Roche) and mounted in a Kopf stereotaxic frame. After exposure of the cranium, five small holes were drilled into the cranium to accommodate surgical screws, one of which served as ground. Another larger hole was drilled over the OFC in the left hemisphere [center of the hole 3.2 mm anterior, 3.2 mm lateral to bregma according to Paxinos and Watson (1996)]. The dura was opened, and the exit bundle of the hyperdrive was lowered onto the exposed cortex. The hole was subsequently filled with a silicone elastomer (Kwik-Sil; World Precision Instruments), after which the hyperdrive was anchored to the screws with dental cement. The hyperdrive consisted of an array of 12 individually drivable tetrodes and two reference electrodes (13 m nichrome wire; Kanthal), spaced apart by at least 310 m (Gray et al., 1995; Gothard et al., 1996). Immediately after surgery, all tetrodes and reference electrodes were advanced 1 mm into the brain; in the course of the next 3 d, the tetrodes were gradually lowered until the OFC was reached. Animals were allowed to recover at least 7 d before the start of the recordings. Before each session started, tetrodes were lowered with increments of 40 m to search for novel neurons to be recorded, after which the animal was brought back to his home cage and left to rest for at least 2 h for the brain tissue to stabilize. Electrophysiological recordings were made using a Cheetah recording system (Neuralynx). Signals from the individual leads of the tetrodes were passed through a low noise unity-gain field-effect transistor preamplifier, insulated multiwire cables, and a 72 channel commutator (Dragonfly) to digitally programmable amplifiers (gain 5000 times; bandpass filtering khz). Amplifier output was digitized at 32 khz to record spike waveforms and stored on a Windows NT station. The occurrence of task events in the behavioral chamber was recorded simultaneously. When all experiments with a given rat were finished, tetrode positions were marked by passing a 10 s, 25 A current through one of the leads of each tetrode. After 24 h, animals were perfused transcardially with a 0.9% saline solution followed by 10% formalin. After removal from the skull, brains were stored in a 10% formalin solution for several days before sectioning. Brain sections of 40 m were cut using a vibratome and were Nissl stained to reconstruct the tracks and final positions of the tetrodes. This showed that recording sites ranged from 2.7 to 4.7 mm anterior to bregma, and were limited to the ventral and lateral orbital regions of the OFC. Recording depth ranged from 3.0 to 5.5 mm (Paxinos and Watson, 1996) (Fig. 1). Data analysis Behavior Behavioral data were analyzed using SPSS for Windows (version 11.0). Unless otherwise stated, results are expressed as mean SEM values. We distinguished two measures to analyze reaction times during the task: movement time was defined as the duration between nose retraction from the odor port and nose entry into the fluid well, and overall response time was defined as the duration of the entire sequence of odor sampling (starting at odor presentation) and nose entry into the fluid

3 8592 J. Neurosci., August 20, (34): van Duuren et al. Population Coding of Reward Magnitude Figure 1. Localization of tetrode recording sites. As indicated by rectangles, recordings in all rats were localized in the ventral and lateral regions of the OFC, between 2.7 and 4.7 mm anterior from bregma. Recording depth ranged from 3 to 5.5 mm (Paxinos and Watson, 1996). Indicated by black arrows in the histological section are several partially visible tetrode tracks. Black asterisks mark the lesion sites that show the final position of three tetrodes. well. The mean response times per reward magnitude were obtained from all corresponding trial types within all the sessions that were used for analysis. These measures were compared across different trial types with the nonparametric Kruskal Wallis test ( p 0.05), followed by a post hoc Mann Whitney U test ( p 0.05). Electrophysiology Isolation of single units. Single units were isolated by off-line cluster cutting procedures (BBClust/MClust 3.0). Before a cluster of spikes was accepted as a single unit, several parameters and graphs were checked visually: the averaged waveform across the four leads, the cluster plots showing spike parameter distributions such as peak amplitudes across the four dimensions, the autocorrelogram, and the spike interval histogram. Because the absence of spike activity during the refractory period (2 ms) is indicative for good isolation, units of which the autocorrelogram and the spike interval histogram revealed activity during this period were removed from the analysis. For the ensemble analysis, all remaining units were taken into consideration. Variability of the representation of reward magnitude. To examine variability in firing rates to reward magnitude, we calculated two variability measures [these have also been used for indicating sparseness of neural coding [cf. Rolls and Tovee (1995) and Perez-Orive et al. (2002)]. Briefly, the population variability (S pop ) is indicative of the variability in the mean firing rate of single cells across the population, regardless of reward size, and is computed by S N r 2 r 2 with r 1 r N 1 r 2 j and r 2 1 r 2 j, (1) N j 1 N j 1 where N indicates the number of units and r j the mean firing rate of neuron j during a particular trial phase, averaged across all three reward conditions. Thus, r represents the mean firing rate in the population and r 2 the mean squared firing rate. The parameter variability (S par ), which is indicative of a single cell s response variability attributable to differences in reward size, was calculated in a similar manner, but index j now indicates each reward size and N the total number of reward sizes (N 3in the current study). Both measures were calculated for the waiting and N N reward delivery phase within a time frame of 1.5 s; for the movement period, the frame was 1.0 s. Values ranged between 0 and 1, with 1 representing the maximum variability attainable. Ensemble analysis: general principles of population coding. Population activity was examined using two different reconstruction methods, the Bayesian method and template matching. Before explaining both methods in more detail, we will first point out the general concepts behind them. A central postulate in systems neurophysiology is that macroscopic parameters (usually in the external world, such as the speed of a moving object) are encoded in the patterns of action potentials generated by neurons. Given these spike patterns, one may also ask how one could make sense out of them, that is, how one could understand what type of information about macroscopic objects or properties they represent. To operationalize this general task, we may ask what a given spike train can tell the experimenter about the stimulus or cognitive process that gave rise to the spike train in the first place (Rieke et al., 1997). For example, given a situation in which a particular spike train may have been evoked by any of three sensory stimuli, the task for a neutral observer, not aware of the actual stimulus, is to decide which of these stimuli was in fact applied to elicit the spike train. This process of reconstructing the original stimulus (or other macroscopic variables) from the spike train is termed decoding. The current objective is to decode the variable reward magnitude from multineuron spike trains. This can be attempted for spike trains generated in anticipation of the reward (expected magnitude, as a cognitive variable), or in relation to the actual reward, as it is delivered to the animal. Decoding can be accomplished successfully on the basis of single-cell firing activity (Bialek et al., 1991). If one would choose to use the average firing rate of a single neuron to decode a macroscopic variable, the procedure would rely on one scalar value. In ensemble recordings studies, however, the firing activity of many neurons offers a potentially much richer source of information; firing-rate values of all simultaneously recorded neurons can be used for decoding, and the series of firing-rate values from all neurons is conceptualized as a population vector. If one records N neurons in a given session, a population vector v can be represented as a series of firing rate values: v (s 1, s 2,...,s N ), where s i is the firing-rate value (scalar) of the first neuron, etc. Geometrically, a vector is an entity in Euclidean space that has both magnitude and direction; for instance, a vector composed of the firing rates of two cells [v (s 1, s 2 )] can be rendered as an arrow in a two-dimensional plane. If a neutral observer is only provided with a sample of firing-rate values of a given set of recorded neurons, but has no knowledge of how macroscopic parameters map onto their firing responses, he would not be able to decode those parameters successfully. Thus, knowledge is needed as to how neurons are tuned to the parameter under study; in other words, we need to know what the standard parameter mapping or encoding of these neurons is. Thus, in addition to needing a population vector for actually decoding the sample set of firing-rate responses back to the value of the macroscopic parameter (i.e., the population vector for decoding), we also should have a template, or separate set of firing-rate values that tells us what the standard mapping from the parameter onto firing-rate responses is (i.e., the population vector for encoding). In the current study, a total of 15 trials for each of the five odor outcome pairs was available, and our standard approach was to extract the template (or encoding vector) from the last six trials of each pair (i.e., trials 10 15), whereas the sample set for decoding was obtained from trials 1 9 of each pair (see below). Given the population vectors for encoding and decoding, some kind of mathematically explicit comparison between the vectors must next be performed to complete the task of decoding. The sample (decoding) set of firing rates must be compared with the template (encoding vector) to reconstruct which macroscopic parameter value was the most likely one giving rise to the observed sample responses. Although both the template-matching and Bayesian reconstruction methods are described below in more detail, the template-matching method works essentially as follows. If we define a template (encoding vector) for two cells as [t (t 1, t 2 )] with t 1 being the firing-rate response of cell 1 to a given parameter, and a to-be-decoded sample of the same cells as [d (d 1, d 2 )], then it is straightforward to visualize these two vectors as two arrows in two-

4 van Duuren et al. Population Coding of Reward Magnitude J. Neurosci., August 20, (34): dimensional space, having the same origin. The similarity (or degree of matching, hence the term template matching ) between the vectors can be expressed as the cosine of the angle between the two vectors, as further explained below. Although population-coding methods as described have the potency to yield important insights bridging gaps between the single-neuron and behavioral-cognitive levels, some limitations can be delineated. The objective of decoding is naturally limited to the parameter that was varied (in this case, reward magnitude); thus the OFC is likely to encode other macroscopic variables [cf. O Doherty et al. (2001), Wallis and Miller (2003), and Roesch et al. (2006)], which, however, should be addressed separately. Second, because we only measure and consider neural correlates of reward in OFC, no conclusions can be drawn as to how OFC populations come to express the capacity for reward-size coding; afferent pathways and intracortical mechanisms for achieving this capacity must be studied separately. Third, whereas in the current study mean firing rate (per trial phase) was used as a measure of neural response, other aspects of firing patterns, such as related to spike timing, may make additional contributions to population coding (cf. Narayanan and Laubach, 2005). In addition to these aspects, there is no code in the neural patterns that could be deciphered, or at least the current methods do not provide a way to do this, if possible at all. Population coding of reward magnitude: template matching. For the main analysis, only the three positively rewarded trial types were taken into consideration, because the quinine and nonreinforced conditions did not yield enough trials for a meaningful analysis except for a control procedure (see Results). As pointed out above, the sessions, all containing 15 correctly performed trials per reward size, were divided into two blocks, of which the first block (trials 1 9) was used for decoding and the second (trials 10 15) for encoding. The first nine trials of the session were chosen for decoding to be able to examine how the representation of reward information builds up during the initial learning phase in the task. We also examined whether a random selection of trials for decoding would provide a similar result; this analysis showed that with randomly selected trials, decoding scores were obtained that were generally higher than with the trials 1 9. For the analysis, two vectors were constructed for each reward size, denominated as x (x 1, x 2,...,x N ) and y ( y 1, y 2,...,y N ), containing the spike counts within a specified time window for the encoding (x) and the decoding (y) block, with x i and y i indicating the spike count of cell i averaged across trials. Thus, the population vector x is used for the encoding part of the procedure (i.e., for determining the template or tuning curves of the cells toward reward magnitude, using the last part of the session, trials 10 15; see above). The population vector y is used for the decoding part of the procedure, in which the spike counts, specific for reward sizes, are taken from the same cells, but now from the first part of the session, trials 1 9). The two vectors were then compared to calculate the decoding score, which is the percentage of correctly identified reward amounts in the decoding phase, based on the activity patterns found in the encoding phase (Fig. 2). Hence, the ensemble code for reward magnitude is made up of the different firing rates of all recorded cells combined in the encoding and decoding phase in relation to reward magnitude. For each trial phase in which we examined population coding of reward size, we used a standard time window, corresponding with the duration of that particular phase within the trial. Reconstruction of reward size for the period during which the animal moved from the sampling port to the fluid well was done with a time frame of 1 s, whereas for the period when the animal awaited reinforcement with its nose in the fluid well, a time frame of 1.5 s was used. During the reward delivery phase, the time frame in which reward size was reconstructed was 10 s, unless otherwise mentioned. With template matching, the similarity between the vectors containing the spike count in the defined time window for the encoding and decoding block was calculated by computing the cosine of the angle between them (Lehky and Sejnowski, 1990; Zhang et al., 1998; Louie and Wilson, 2001). A value of 1 represents an exact similarity between the two vectors and 1 the exact opposite, whereas 0 (i.e., orthogonal) indicates no Figure 2. Schematic representation of the template-matching procedure. This example shows the decoding of 150 l of sucrose solution. First three encoding vectors are generated, one for each reward size, which contain the firing rates of all cells obtained from trials These vectors serve as templates. The sample set (or decoding vector) contains the firing rates obtained in trials 1 9. The decoding set is compared with the template to reconstruct which parameter value most likely gave rise to the observed sample response. To this end, the similarity between the encoding and decoding vectors is calculated for all three reward amounts by computing the cosine of the angle ( ) between the decoding vector and each of the three encoding vectors (see also Eq. 3). The highest cosine value max (cos ), which indicates the highest similarity between two vectors, is selected as reconstructed amount of reward. In this example, the decoding vector shows the highest similarity with the encoding vector constructed for 150 l. Hence, 150 l is selected as the reconstructed amount of reward. similarity between the two vectors. We first calculated the inner product of x and y: N i 1 x i y i, (2) where x i and y i indicate the average firing rate of neuron i from a total of N cells within the specified time window for the encoding and decoding block, respectively (Fig. 2). Then the cosine was calculated by N i 1 x i y i cos x y, (3) with the denominator representing the product of the absolute vector lengths. If the decoding spike vector belonging to a particular reward amount produced the highest cosine value with respect to the encoding vector, then that particular reward size was selected as reconstructed amount of reward (Fig. 2). In several graphs, the decoding score (i.e., the percentage of trials in which the amount of reward was correctly reconstructed) was expressed as a function of time or as a function of the size of the reconstruction ensemble, i.e., the group of neurons that was subsampled from the entire population and used for the calculations. The maximum size of the reconstruction ensemble was 37: this value represents the median value of the number of cells recorded in all sessions, which ranged between 26 and 60. Unless otherwise mentioned, a reconstruction ensemble of 37 neurons was used for calculations. Calculations were made for each recording session, after which data were averaged. For the assessment of decoding as a function of size of the reconstruction ensemble, the decoding score was calculated 100 times for each group size, each time with neurons randomly picked from the population recorded in that particular session. Decoding as a function of time was calculated with a reconstruction ensemble of 37 neurons as well: the decoding score was calculated 100 times per time window, each time with randomly picked neurons. The decoding curves were further analyzed by applying linear regression analysis ( p 0.05) and a one-way ANOVA test with Bonferroni correction ( p 0.05). Population coding of reward magnitude: Bayesian reconstruction. Bayesian or probabilistic reconstruction was used as previously described by

5 8594 J. Neurosci., August 20, (34): van Duuren et al. Population Coding of Reward Magnitude Földiák (1993), Sanger (1996), Zhang et al. (1998), and Thiel et al. (2007). Briefly, population vectors were calculated as pointed out above, and decoding was based on the equation of conditional probability: P s y P y P y s P s, (4) where P(s y) indicates the probability of a reward size s given the multineuron spike pattern vector y. P(s) indicates the prior probability of reward size, which does not need to be calculated; it has a fixed value of 1/3 because of the three different amounts of rewards that were applied with equal probability across trials. The probability P(y) for the spikecontaining decoding vector y to occur does not need to be calculated either: because the reconstructed amount of reward was the most probable reward size of the probability distribution (see below), this can be considered as an unnecessary scaling factor. Under the assumptions of a Poisson distribution of spike timing and cells firing independently from each other, P(y s) for every reward size s was computed by N N x i s yi P y s P y x s P yi x s exp x i 1 i 1 y i! i s, (5) where x i is the product of the length of the time window and the average firing rate f i (s) of cell i for a given reward size, and x(s) is the multineuron spike pattern vector used for encoding (i.e., determining the tuning of neurons to reward size). We calculated the logarithms of the probabilities P(y i x(s)) to avoid working with extremely small values. The most probable reward size of the probability distribution was taken as the reconstructed amount of reward, which is the maximum of P(y x(s)). The latter quantity is equivalent to P(s y) (see Eq. 4). Further data analysis was done as described for template matching. Analyzing the data with these different decoding methods revealed that in all cases decoding scores for reward amount were slightly higher with template matching than with the Bayesian method. To elaborate on the lower decoding scores for Bayesian method as observed in our results, we recalculated decoding with various variants of the Bayesian method. When the spike timing distributions are approximated by a Poisson distribution, the firing rates within the decoding vector are rounded to integers (see Eq. 5). To examine whether the use of integer values contributes to a lower performance, we tried a gamma distribution instead, but did not observe an improvement. When a Gaussian was used as distribution of spike timing, decoding scores were comparable with those obtained under a Poisson distribution as well. This may be because of the high proportion of cells with low firing rates, which causes the use of a Gaussian distribution to be inappropriate. It should be noted, however, that cells with low firing rates were found to be important for reconstruction, because their removal from the reconstruction ensemble led to lower decoding scores than when these cells were taken into account. Results Behavior Data from 17 recording sessions were used for the analysis, obtained from six rats. For all three positively rewarded trial types, animals needed on average 17 trials to reach the criterion of 15 successful trials per reward size (0.05 ml, , proportion correct 88%; 0.15 ml, , proportion correct 88%; 0.30 ml, , proportion correct 83%). For the nonreinforced and quinine trial types, animals made on average and go responses, respectively (sucrose vs quinine or vs nonreinforced, p 0.000; paired sampled t test). The reason why rats initially perform responses during quinine trials might be that they have to taste the reinforcing fluid to learn which outcome to avoid or to approach based on the initially neutral olfactory stimuli. Examination of the movement time showed no significant differences between the positively reinforced trial types (0.05 ml, s; 0.15 ml, s; 0.30 ml, s; number of trials per reward size, 255). When the positively reinforced trial types were combined ( s), movement time for sucrose trials was significantly shorter compared with quinine trials, but not compared with the nonrewarded trial types (quinine go responses, s; n 87; nonrewarded go responses, s; n 88). The overall response time revealed that animals responded significantly faster to obtain the highest amount of reward compared with the lowest amount; comparison of these reward amounts with the middle-sized reward showed no significant difference (0.05 ml, s; 0.15 ml, s; 0.30 ml, s). Thus, learning within this task was evident from the faster overall response time for the largest reward and from the lower number of quinine responses, which were performed with a slower movement time as well compared with responses during sucrose trials. Electrophysiology Variability in the representation of reward magnitude Over the course of 17 sessions recorded in six animals, a total number of 683 single units was obtained. The number of single units recorded per session ranged from 26 to 60 (mean SEM, ), with a mean firing rate of spikes/s. Neurons could be specifically activated during several task phases, including odor sampling, the behavioral phase in which animals moved from the odor sampling port to the fluid well ( movement phase ), the period of waiting for reinforcement in the fluid well, and the reward period (i.e., the period after reward delivery lasting for 10 s) (Fig. 3). Examination of ensemble activity focused mainly on two task periods, namely the waiting and reward delivery periods, because we expected reward predictive information to be coded especially during the waiting period, and information about the size of the actual reward after reward delivery. When of particular interest, population coding was also examined for the movement phase. Because the task was not designed to determine whether differential responses during cue sampling were attributable to different sensory inputs (odor identity) or expectancy of varying reward magnitude, results regarding the coding of expected reward magnitude during this phase would be inconclusive. Hence, ensemble activity in this task period was not examined. Before studying population coding principles themselves, we first aimed to assess (1) to what extent the firing-rate variability of single cells is attributable to effects of reward size and (2) whether whole OFC populations exhibit a high or low degree of overall firing-rate variability, regardless of reward size. This was done by calculating two variability measures, namely S par and S pop. The mean S par, indicating the response variability of individual neurons associated with variations in reward size (their tuning curves), was 0.23 for the waiting period and 0.28 for the reward delivery phase (Fig. 4A,B). The mean S pop, which represents the variability in firing rate across the population regardless of reward size effects, was 0.70 for the waiting period and 0.73 for the reward period (Fig. 4C,D). Similar findings were obtained for the movement period, namely a mean S par value of 0.25 and a mean S pop value of These results show, first, that there is a high variability of firing rates throughout the population, as illustrated by the high S pop values throughout the three trial phases. Second, this phenomenon is not matched by a similarly high rewardrelated variability when individual cells are considered: OFC neurons displayed less variability in their tuning curves to reward size, because for all three task periods the large majority of the cells had values 0.5 (movement period, 76%; n 518; waiting period, 79%; n 541; reward period, 76%; n 520). Apparently, other factors must be present in addition to the reward-

6 van Duuren et al. Population Coding of Reward Magnitude J. Neurosci., August 20, (34): Population coding of expected and actual reward magnitude We will describe first an application of the Bayesian method and template matching in which the magnitude of reward was decoded from ensemble activity in the positively rewarded trial types across the first nine trials of the session, whereas trials were used for encoding. During the movement, waiting, and reward periods, the magnitude of reward was decoded from the population activity with a percentage of correct performance significantly above the 1/3 chance level for both methods (oneway ANOVA; p in all three cases). When the decoding score was plotted as a function of the number of cells participating in the reconstruction ensemble, regression analysis showed that with template matching the slope of the decoding curve was significantly positive for all three task phases ( p in all cases), meaning that the decoding performance improved with an increasing amount of cells (Fig. 5). The maximum decoding scores were 57% (at n 30 cells) for the movement period, and 48% (at n 36 cells) and 76% (n 36 cells) for the waiting and reward periods, respectively. We also calculated the decoding score for the subgroup of the sessions that did not show a significant difference in response latency between the lowest and highest amount of reward. Reward magnitude could be reconstructed above chance level in this group as well. This finding indicates that significant population coding is present in OFC even in the absence of overt behavioral differences, which corroborates the notion that OFC activity is not necessarily tied to motor behavior and involves a cognitive process. The Bayesian method produced similar although generally somewhat lower decoding scores, with maximum scores of 44% for the movement period and waiting period (at, respectively, n 26 and n 24 cells), and 79% for the reward period (at n 36 cells). Also for this method, regression analysis indicated a significantly rising score with increasing ensemble size for the reward period ( p 0.000). For the waiting and movement period, however, no significant effect was found. Across all analyses presented here, Bayesian reconstruction and template matching produced similar results for the reward period, whereas for the waiting-anticipation phase the decoding scores were similar or lower for the Bayesian compared with template-matching method. Such differences may be attributable to several factors (see Materials and Methods). Below, we will concentrate on results obtained with template matching. Figure3. Summaryoftask-relatedfiringactivityof32simultaneouslyrecordedOFCneurons during a single session. The horizontal axis denotes time (in seconds), and the vertical axis denotes cell number. Color-coded firing rates were calculated by averaging activity over all rewarded trial types during the session, with bin sizes of 100 ms. A C, Neural activity is synchronizedonodorpresentation(a), theendoftheodorpoke(b), andrewarddelivery(c). Onset of the waiting period is 1.5 s before reward delivery. Colors indicate firing rates that are normalized relative to the maximum firing rate of each neuron, according to a hotness scale shown on the right. The reward-size variations in firing activity of these cells are used as the basis for studying the ensemble code of reward magnitude. modulated response profiles of the individual cells to explain the high variability in the overall population, e.g., differences in baseline firing rate or differences in responsivity during various task phases, although a small subgroup was present with a very high parameter variability (range, ) in all three task phases. In conclusion, whereas firing rates throughout OFC populations are highly variable, modulation of single-cell firing rate that is attributable to reward size is relatively subtle, in line with a previous analysis of single-unit activity in OFC (van Duuren et al., 2007). Temporal resolution of ensemble coding Because the results indicated reward magnitude to be represented in population activity during task performance in a manner that was consistent across decoding and encoding blocks, the question arose what the temporal resolution of population coding was. This question is relevant for understanding OFC function, because when significant coding is present across short time intervals, this may facilitate rapid decision making and fast behavioral responding. We first asked how the decoding score depends on the width of the time window used for encoding and decoding applied to either the waiting or reward period. To this end, we selected smaller time windows from the overall time windows of the waiting and reward period (1.5 and 10 s, respectively), and calculated the decoding score for time windows of increasing duration. This provides a cumulative measure of the percentage of decoded reward size over time within each trial phase. The waiting period was divided into six time windows that started at 250 ms and went up to 1.5 s, with increments of 250 ms. Time windows used for the reward period were, in addition to the six windows of 250 ms, 2, 3, 4, 5, 6, 8, and 10 s. Regression analysis

7 8596 J. Neurosci., August 20, (34): van Duuren et al. Population Coding of Reward Magnitude obtained with a decoding window of 1.5 s was significantly above chance level ( p 0.001), and furthermore revealed that there was no significant variation between the various window widths. In contrast, during the reward period, all time windows with the exception of the first one differed significantly from chance level ( p 0.000). Furthermore, the first two time windows (i.e., 250 and 500 ms) had a significantly lower decoding score than the final six time windows (i.e., 3, 4, 5, 6, 8, and 10 s). Thus, for a significant success rate in decoding actual reward size, only a short time segment ( 0.5 s) of the population response is needed, whereas the time segment needed for a significant success rate in decoding expected reward is somewhat longer, namely 1.5 s. We next asked how the decoding capacity of the population varies over time across consecutive segments of the waiting and reward periods. This question probes Figure 4. Distribution of population variability and parameter variability across orbitofrontal cell populations. A, C, Waiting the time resolution at which significant period; B, D, reward period. Values of population variability across sessions were generally in a high range between 0.4 and 1.0, reward-predictive information is represented across consecutive time segments with means of 0.71 and 0.73 for the waiting and reward periods, respectively. Parameter variability values varied more strongly, spanning the whole range from 0.0 to 1.0 for both periods, with means of 0.23 and 0.28 for the waiting and reward periods, of a behavioral task phase. Therefore the respectively. Similar findings were obtained for the movement period, namely a mean S par value of 0.25 and a mean S pop value of percentage of correctly decoded reward 0.76 (data not shown). Note the subgroup of cells with very high parameter variability ( ). amount was calculated for successive time segments across the entire length of the overall time windows of the waiting and reward period. Thus, a sequential method was adopted instead of the cumulative method used above. For the waiting period, we used three consecutive time segments of 500 ms; for the reward period, decoding was calculated during the first 2sintime segments of 500 ms, followed by four time segments of 1 s, and the final two segments of 2 s. During both the waiting and reward period, regression analysis and one-way ANOVA did not show a significant variation in decoding success, indicating that the amount of decodable information did not significantly differ across the various time segments within this task period (Fig. 6C,D). Figure 5. Decoding of reward magnitude with template matching: dependence on ensemble size. The horizontal axis indicates the size of the reconstruction ensemble, and the vertical axis indicates the percentage of trials in which reward size was correctly decoded. In the graphs presented here and in Figures 6 7 and 9 10, the horizontal dashed line indicates chance level (33.3%), and dotted lines flanking the curves represent the 95% confidence interval(two times the SE of proportion). indicated that with template matching, the percentage of reward amount correctly decoded during the waiting and reward period rose with increasing length of the decoding window, with the slope of the decoding curve being significantly positive (waiting period, p 0.039; reward period, p 0.000) (Fig. 6A,B). The maximum decoding scores obtained were 47 and 80% for the waiting and reward periods, respectively. A one-way ANOVA indicated that during the waiting period, only the decoding score Development of population coding during task progression Because the data on response times and the number of trials per reinforcer indicated that in the course of a session animals generated predictions about response outcome and learned to discriminate between different amounts of reward, we addressed the question how the representation of reward magnitude by population activity evolved in the course of learning during a session. To this end, the decoding block was divided into four consecutive two-trial blocks (block 1, trials 1 and 2; block 2, trials 3 and 4, block 3, trials 5 and 6; block 4, trials 7 and 8), and decoding was compared between these four blocks. Trial numbers were used for encoding (as above). In all four trial blocks and for both the waiting and reward period, reward magnitude was decoded from the population activity above chance level (one-way ANOVA; p 0.000). Regression analysis indicated for the waiting period a significant slope of the decoding score for blocks 2, 3, and 4. Decoding with the first trial block differed significantly from all three other blocks, with an improvement in decoding score as the task progressed, except for the final (fourth) trial block. The second block did not differ from the third block, whereas the third block differed significantly from the fourth

8 van Duuren et al. Population Coding of Reward Magnitude J. Neurosci., August 20, (34): (Fig. 7A). During the reward period, all blocks showed a significant slope of the decoding curves. Furthermore, the decoding curves differed significantly between the first and the second blocks and the third versus the fourth, with an improvement of decoding success; the second block did not differ from the third block (Fig. 7B). An additional two-way ANOVA indicated no interaction between cell group size and trial block. The trial-block sequence observed for the waiting period, in which block 1 is flat, whereas the blocks 2, 3, and 4 show increasing decoding with an augmenting number of cells as indicated by the regression analysis, suggests that OFC ensembles have only a weak predictive ability at the beginning of the session (the first block), which is already consistently present in small subsets of neurons. At this time in the session, the decoding score does not rise with an increasing amount of cells; adding more cells to the ensemble adds both confirmatory and conflicting evidence about the predicted reward size. However, the ability to predict reward size increases with experience over trial blocks (blocks 2, 3, and 4), and the decoding score now comes to depend on the number of cells that participate in the ensemble, as indicated by the regression analysis. This means that at this point, the code has developed in a more consistent and redundant manner within OFC, with less conflicting neural evidence. Figure 6. Decoding of reward magnitude within specific trial phases. The size of the reconstruction ensemble was 37 neurons. A, B, Decoding score using time windows of increasing duration for the waiting (A) and the reward delivery (B) periods. The horizontal axis shows the size of the time window (in seconds) from which spikes were taken for reconstruction, and the vertical axis shows the percentage of correctly decoded trials. C, D, The decoding success for consecutive temporal segments in the waiting and reward periods, respectively. The horizontal axis now shows the time segment of trial phase used for reconstruction. In C, the decoding segmentsaretheintervals(0 0.5),( ), and( ) srelativetofluidpokeonset. InD, the segments are (0 0.5), ( ), ( ), ( ), ( ), ( ), ( ), ( ), ( ), and ( ) s relative to reward delivery. Nonspecific drift across trials as a possible confounding factor The increased decoding success found for later trial blocks relative to earlier ones may be explained either by a learning effect or by a nonspecific drift in ensemble responses over trials, because the final block of five trials (i.e., trials 10 15) was used for the encoding. To check whether the latter, confounding possibility would apply, we compared the two-trial decoding blocks when the encoding block was situated at the session end (trials 10 15) with two-trial blocks when the encoding block was situated at the start of the session (encoding block, trials 1 6; decoding block 1, trials 7 and 8; block 2, trials 9 and 10, block 3, trials 11 and 12; block 4, trials 13 and 14). Because outcome-prediction learning is expected to be unevenly distributed across trials (with a rapid decrease in go-responses for the quinine and nonrewarded trials early in the session), it is predicted that these decoding procedures should yield separate or at most partially overlapping curves for the waiting period if the progressive change in decoding is indeed attributable to learning. Thus, the curve with encoding by late trials should rise more steeply and then stabilize compared with its temporal mirror image. As illustrated in Figure 7C, showing the average of the decoding curves per trial block for the waiting period (as in Fig. 7A), the curves indeed confirmed this prediction. A one-way ANOVA ( p 0.05) indicated that for the waiting period all trial blocks of the two different encoding conditions differed significantly from each other, except for the fourth block. For the reward period, in which additional learning may or may not take place, the first and the fourth trial block differed significantly between the two encoding conditions, with a steeper rise across trial blocks when the encoding block was at the session end (Fig. 7D). Altogether, these results indicate that progressive learning of odor reward associations coupled to motor responses is accompanied by a quick rise in population coding of expected reward magnitude during the waiting period, followed by relative stabilization: a time course not attributable to nonspecific drift. An additional learning-related increase in population coding appears to take place in the reward period, although this effect is less strong.

Supplementary materials for: Executive control processes underlying multi- item working memory

Supplementary materials for: Executive control processes underlying multi- item working memory Supplementary materials for: Executive control processes underlying multi- item working memory Antonio H. Lara & Jonathan D. Wallis Supplementary Figure 1 Supplementary Figure 1. Behavioral measures of

More information

Introduction to Computational Neuroscience

Introduction to Computational Neuroscience Introduction to Computational Neuroscience Lecture 5: Data analysis II Lesson Title 1 Introduction 2 Structure and Function of the NS 3 Windows to the Brain 4 Data analysis 5 Data analysis II 6 Single

More information

Supplementary Figure 1. Example of an amygdala neuron whose activity reflects value during the visual stimulus interval. This cell responded more

Supplementary Figure 1. Example of an amygdala neuron whose activity reflects value during the visual stimulus interval. This cell responded more 1 Supplementary Figure 1. Example of an amygdala neuron whose activity reflects value during the visual stimulus interval. This cell responded more strongly when an image was negative than when the same

More information

Double dissociation of value computations in orbitofrontal and anterior cingulate neurons

Double dissociation of value computations in orbitofrontal and anterior cingulate neurons Supplementary Information for: Double dissociation of value computations in orbitofrontal and anterior cingulate neurons Steven W. Kennerley, Timothy E. J. Behrens & Jonathan D. Wallis Content list: Supplementary

More information

The individual animals, the basic design of the experiments and the electrophysiological

The individual animals, the basic design of the experiments and the electrophysiological SUPPORTING ONLINE MATERIAL Material and Methods The individual animals, the basic design of the experiments and the electrophysiological techniques for extracellularly recording from dopamine neurons were

More information

Behavioral generalization

Behavioral generalization Supplementary Figure 1 Behavioral generalization. a. Behavioral generalization curves in four Individual sessions. Shown is the conditioned response (CR, mean ± SEM), as a function of absolute (main) or

More information

Supplementary Figure 1 Information on transgenic mouse models and their recording and optogenetic equipment. (a) 108 (b-c) (d) (e) (f) (g)

Supplementary Figure 1 Information on transgenic mouse models and their recording and optogenetic equipment. (a) 108 (b-c) (d) (e) (f) (g) Supplementary Figure 1 Information on transgenic mouse models and their recording and optogenetic equipment. (a) In four mice, cre-dependent expression of the hyperpolarizing opsin Arch in pyramidal cells

More information

Analysis of in-vivo extracellular recordings. Ryan Morrill Bootcamp 9/10/2014

Analysis of in-vivo extracellular recordings. Ryan Morrill Bootcamp 9/10/2014 Analysis of in-vivo extracellular recordings Ryan Morrill Bootcamp 9/10/2014 Goals for the lecture Be able to: Conceptually understand some of the analysis and jargon encountered in a typical (sensory)

More information

Sum of Neurally Distinct Stimulus- and Task-Related Components.

Sum of Neurally Distinct Stimulus- and Task-Related Components. SUPPLEMENTARY MATERIAL for Cardoso et al. 22 The Neuroimaging Signal is a Linear Sum of Neurally Distinct Stimulus- and Task-Related Components. : Appendix: Homogeneous Linear ( Null ) and Modified Linear

More information

SI Appendix: Full Description of Data Analysis Methods. Results from data analyses were expressed as mean ± standard error of the mean.

SI Appendix: Full Description of Data Analysis Methods. Results from data analyses were expressed as mean ± standard error of the mean. SI Appendix: Full Description of Data Analysis Methods Behavioral Data Analysis Results from data analyses were expressed as mean ± standard error of the mean. Analyses of behavioral data were performed

More information

Nature Neuroscience: doi: /nn Supplementary Figure 1. Behavioral training.

Nature Neuroscience: doi: /nn Supplementary Figure 1. Behavioral training. Supplementary Figure 1 Behavioral training. a, Mazes used for behavioral training. Asterisks indicate reward location. Only some example mazes are shown (for example, right choice and not left choice maze

More information

Chapter 6. The Journal of Pharmacology and Experimental Therapeutics 323:61 69, *these authors contributed equally to this article.

Chapter 6. The Journal of Pharmacology and Experimental Therapeutics 323:61 69, *these authors contributed equally to this article. Chapter 6 Pharmacological manipulation of neuronal ensemble activity by reverse microdialysis in freely moving rats: a comparative study of the effects of tetrodotoxin, lidocaine, and muscimol Esther van

More information

Summary of behavioral performances for mice in imaging experiments.

Summary of behavioral performances for mice in imaging experiments. Supplementary Figure 1 Summary of behavioral performances for mice in imaging experiments. (a) Task performance for mice during M2 imaging experiments. Open triangles, individual experiments. Filled triangles,

More information

Supplementary Figure 1. Recording sites.

Supplementary Figure 1. Recording sites. Supplementary Figure 1 Recording sites. (a, b) Schematic of recording locations for mice used in the variable-reward task (a, n = 5) and the variable-expectation task (b, n = 5). RN, red nucleus. SNc,

More information

Brain and Cognitive Sciences 9.96 Experimental Methods of Tetrode Array Neurophysiology IAP 2001

Brain and Cognitive Sciences 9.96 Experimental Methods of Tetrode Array Neurophysiology IAP 2001 Brain and Cognitive Sciences 9.96 Experimental Methods of Tetrode Array Neurophysiology IAP 2001 An Investigation into the Mechanisms of Memory through Hippocampal Microstimulation In rodents, the hippocampus

More information

Microcircuitry coordination of cortical motor information in self-initiation of voluntary movements

Microcircuitry coordination of cortical motor information in self-initiation of voluntary movements Y. Isomura et al. 1 Microcircuitry coordination of cortical motor information in self-initiation of voluntary movements Yoshikazu Isomura, Rie Harukuni, Takashi Takekawa, Hidenori Aizawa & Tomoki Fukai

More information

SUPPLEMENTARY INFORMATION. Table 1 Patient characteristics Preoperative. language testing

SUPPLEMENTARY INFORMATION. Table 1 Patient characteristics Preoperative. language testing Categorical Speech Representation in the Human Superior Temporal Gyrus Edward F. Chang, Jochem W. Rieger, Keith D. Johnson, Mitchel S. Berger, Nicholas M. Barbaro, Robert T. Knight SUPPLEMENTARY INFORMATION

More information

Within-event learning contributes to value transfer in simultaneous instrumental discriminations by pigeons

Within-event learning contributes to value transfer in simultaneous instrumental discriminations by pigeons Animal Learning & Behavior 1999, 27 (2), 206-210 Within-event learning contributes to value transfer in simultaneous instrumental discriminations by pigeons BRIGETTE R. DORRANCE and THOMAS R. ZENTALL University

More information

Birds' Judgments of Number and Quantity

Birds' Judgments of Number and Quantity Entire Set of Printable Figures For Birds' Judgments of Number and Quantity Emmerton Figure 1. Figure 2. Examples of novel transfer stimuli in an experiment reported in Emmerton & Delius (1993). Paired

More information

Nature Neuroscience: doi: /nn Supplementary Figure 1. Trial structure for go/no-go behavior

Nature Neuroscience: doi: /nn Supplementary Figure 1. Trial structure for go/no-go behavior Supplementary Figure 1 Trial structure for go/no-go behavior a, Overall timeline of experiments. Day 1: A1 mapping, injection of AAV1-SYN-GCAMP6s, cranial window and headpost implantation. Water restriction

More information

Error Detection based on neural signals

Error Detection based on neural signals Error Detection based on neural signals Nir Even- Chen and Igor Berman, Electrical Engineering, Stanford Introduction Brain computer interface (BCI) is a direct communication pathway between the brain

More information

Theta sequences are essential for internally generated hippocampal firing fields.

Theta sequences are essential for internally generated hippocampal firing fields. Theta sequences are essential for internally generated hippocampal firing fields. Yingxue Wang, Sandro Romani, Brian Lustig, Anthony Leonardo, Eva Pastalkova Supplementary Materials Supplementary Modeling

More information

Supplementary Figure 1

Supplementary Figure 1 Supplementary Figure 1 Miniature microdrive, spike sorting and sleep stage detection. a, A movable recording probe with 8-tetrodes (32-channels). It weighs ~1g. b, A mouse implanted with 8 tetrodes in

More information

Electrophysiological and firing properties of neurons: categorizing soloists and choristers in primary visual cortex

Electrophysiological and firing properties of neurons: categorizing soloists and choristers in primary visual cortex *Manuscript Click here to download Manuscript: Manuscript revised.docx Click here to view linked Referenc Electrophysiological and firing properties of neurons: categorizing soloists and choristers in

More information

Supplementary Material for

Supplementary Material for Supplementary Material for Selective neuronal lapses precede human cognitive lapses following sleep deprivation Supplementary Table 1. Data acquisition details Session Patient Brain regions monitored Time

More information

Single cell tuning curves vs population response. Encoding: Summary. Overview of the visual cortex. Overview of the visual cortex

Single cell tuning curves vs population response. Encoding: Summary. Overview of the visual cortex. Overview of the visual cortex Encoding: Summary Spikes are the important signals in the brain. What is still debated is the code: number of spikes, exact spike timing, temporal relationship between neurons activities? Single cell tuning

More information

Representation of negative motivational value in the primate

Representation of negative motivational value in the primate Representation of negative motivational value in the primate lateral habenula Masayuki Matsumoto & Okihide Hikosaka Supplementary Figure 1 Anticipatory licking and blinking. (a, b) The average normalized

More information

Ch.20 Dynamic Cue Combination in Distributional Population Code Networks. Ka Yeon Kim Biopsychology

Ch.20 Dynamic Cue Combination in Distributional Population Code Networks. Ka Yeon Kim Biopsychology Ch.20 Dynamic Cue Combination in Distributional Population Code Networks Ka Yeon Kim Biopsychology Applying the coding scheme to dynamic cue combination (Experiment, Kording&Wolpert,2004) Dynamic sensorymotor

More information

Are Retrievals from Long-Term Memory Interruptible?

Are Retrievals from Long-Term Memory Interruptible? Are Retrievals from Long-Term Memory Interruptible? Michael D. Byrne byrne@acm.org Department of Psychology Rice University Houston, TX 77251 Abstract Many simple performance parameters about human memory

More information

Nature Methods: doi: /nmeth Supplementary Figure 1. Activity in turtle dorsal cortex is sparse.

Nature Methods: doi: /nmeth Supplementary Figure 1. Activity in turtle dorsal cortex is sparse. Supplementary Figure 1 Activity in turtle dorsal cortex is sparse. a. Probability distribution of firing rates across the population (notice log scale) in our data. The range of firing rates is wide but

More information

Neuronal selectivity, population sparseness, and ergodicity in the inferior temporal visual cortex

Neuronal selectivity, population sparseness, and ergodicity in the inferior temporal visual cortex Biol Cybern (2007) 96:547 560 DOI 10.1007/s00422-007-0149-1 ORIGINAL PAPER Neuronal selectivity, population sparseness, and ergodicity in the inferior temporal visual cortex Leonardo Franco Edmund T. Rolls

More information

Intelligent Adaption Across Space

Intelligent Adaption Across Space Intelligent Adaption Across Space Honors Thesis Paper by Garlen Yu Systems Neuroscience Department of Cognitive Science University of California, San Diego La Jolla, CA 92092 Advisor: Douglas Nitz Abstract

More information

Neural Correlates of Olfactory Recognition Memory in the Rat Orbitofrontal Cortex

Neural Correlates of Olfactory Recognition Memory in the Rat Orbitofrontal Cortex The Journal of Neuroscience, November 1, 2000, 20(21):8199 8208 Neural Correlates of Olfactory Recognition Memory in the Rat Orbitofrontal Cortex Seth J. Ramus and Howard Eichenbaum Department of Psychology,

More information

Introduction to Computational Neuroscience

Introduction to Computational Neuroscience Introduction to Computational Neuroscience Lecture 11: Attention & Decision making Lesson Title 1 Introduction 2 Structure and Function of the NS 3 Windows to the Brain 4 Data analysis 5 Data analysis

More information

The effect of sample duration and cue on a double temporal discrimination q

The effect of sample duration and cue on a double temporal discrimination q Available online at www.sciencedirect.com Learning and Motivation 39 (2008) 71 94 www.elsevier.com/locate/l&m The effect of sample duration and cue on a double temporal discrimination q Luís Oliveira,

More information

Spike Sorting and Behavioral analysis software

Spike Sorting and Behavioral analysis software Spike Sorting and Behavioral analysis software Ajinkya Kokate Department of Computational Science University of California, San Diego La Jolla, CA 92092 akokate@ucsd.edu December 14, 2012 Abstract In this

More information

Supplemental Information. A Visual-Cue-Dependent Memory Circuit. for Place Navigation

Supplemental Information. A Visual-Cue-Dependent Memory Circuit. for Place Navigation Neuron, Volume 99 Supplemental Information A Visual-Cue-Dependent Memory Circuit for Place Navigation Han Qin, Ling Fu, Bo Hu, Xiang Liao, Jian Lu, Wenjing He, Shanshan Liang, Kuan Zhang, Ruijie Li, Jiwei

More information

Supporting Information

Supporting Information 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 Supporting Information Variances and biases of absolute distributions were larger in the 2-line

More information

SpikerBox Neural Engineering Workshop

SpikerBox Neural Engineering Workshop SpikerBox Neural Engineering Workshop A Workshop Curriculum for Grades 9-12 Developed as a two-day, two hours/day workshop Developed by UW Graduate Students: Stephanie Seeman, Bethany Kondiles, and Katherine

More information

DATA MANAGEMENT & TYPES OF ANALYSES OFTEN USED. Dennis L. Molfese University of Nebraska - Lincoln

DATA MANAGEMENT & TYPES OF ANALYSES OFTEN USED. Dennis L. Molfese University of Nebraska - Lincoln DATA MANAGEMENT & TYPES OF ANALYSES OFTEN USED Dennis L. Molfese University of Nebraska - Lincoln 1 DATA MANAGEMENT Backups Storage Identification Analyses 2 Data Analysis Pre-processing Statistical Analysis

More information

The mammalian cochlea possesses two classes of afferent neurons and two classes of efferent neurons.

The mammalian cochlea possesses two classes of afferent neurons and two classes of efferent neurons. 1 2 The mammalian cochlea possesses two classes of afferent neurons and two classes of efferent neurons. Type I afferents contact single inner hair cells to provide acoustic analysis as we know it. Type

More information

brain valuation & behavior

brain valuation & behavior brain valuation & behavior 9 Rangel, A, et al. (2008) Nature Neuroscience Reviews Vol 9 Stages in decision making process Problem is represented in the brain Brain evaluates the options Action is selected

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION doi:10.1038/nature10776 Supplementary Information 1: Influence of inhibition among blns on STDP of KC-bLN synapses (simulations and schematics). Unconstrained STDP drives network activity to saturation

More information

Neural Coding. Computing and the Brain. How Is Information Coded in Networks of Spiking Neurons?

Neural Coding. Computing and the Brain. How Is Information Coded in Networks of Spiking Neurons? Neural Coding Computing and the Brain How Is Information Coded in Networks of Spiking Neurons? Coding in spike (AP) sequences from individual neurons Coding in activity of a population of neurons Spring

More information

Orbitofrontal cortical activity during repeated free choice

Orbitofrontal cortical activity during repeated free choice J Neurophysiol 107: 3246 3255, 2012. First published March 14, 2012; doi:10.1152/jn.00690.2010. Orbitofrontal cortical activity during repeated free choice Michael Campos, Kari Koppitch, Richard A. Andersen,

More information

A Neural Model of Context Dependent Decision Making in the Prefrontal Cortex

A Neural Model of Context Dependent Decision Making in the Prefrontal Cortex A Neural Model of Context Dependent Decision Making in the Prefrontal Cortex Sugandha Sharma (s72sharm@uwaterloo.ca) Brent J. Komer (bjkomer@uwaterloo.ca) Terrence C. Stewart (tcstewar@uwaterloo.ca) Chris

More information

Supporting Information

Supporting Information Supporting Information ten Oever and Sack 10.1073/pnas.1517519112 SI Materials and Methods Experiment 1. Participants. A total of 20 participants (9 male; age range 18 32 y; mean age 25 y) participated

More information

Information Processing During Transient Responses in the Crayfish Visual System

Information Processing During Transient Responses in the Crayfish Visual System Information Processing During Transient Responses in the Crayfish Visual System Christopher J. Rozell, Don. H. Johnson and Raymon M. Glantz Department of Electrical & Computer Engineering Department of

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION doi:10.1038/nature11239 Introduction The first Supplementary Figure shows additional regions of fmri activation evoked by the task. The second, sixth, and eighth shows an alternative way of analyzing reaction

More information

Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011

Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011 Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011 I. Purpose Drawing from the profile development of the QIBA-fMRI Technical Committee,

More information

Data and Statistics 101: Key Concepts in the Collection, Analysis, and Application of Child Welfare Data

Data and Statistics 101: Key Concepts in the Collection, Analysis, and Application of Child Welfare Data TECHNICAL REPORT Data and Statistics 101: Key Concepts in the Collection, Analysis, and Application of Child Welfare Data CONTENTS Executive Summary...1 Introduction...2 Overview of Data Analysis Concepts...2

More information

Spontaneous Cortical Activity Reveals Hallmarks of an Optimal Internal Model of the Environment. Berkes, Orban, Lengyel, Fiser.

Spontaneous Cortical Activity Reveals Hallmarks of an Optimal Internal Model of the Environment. Berkes, Orban, Lengyel, Fiser. Statistically optimal perception and learning: from behavior to neural representations. Fiser, Berkes, Orban & Lengyel Trends in Cognitive Sciences (2010) Spontaneous Cortical Activity Reveals Hallmarks

More information

Stimulus-driven population activity patterns in macaque primary visual cortex

Stimulus-driven population activity patterns in macaque primary visual cortex Stimulus-driven population activity patterns in macaque primary visual cortex Benjamin R. Cowley a,, Matthew A. Smith b, Adam Kohn c, Byron M. Yu d, a Machine Learning Department, Carnegie Mellon University,

More information

Millisecond-scale differences in neural activity in auditory cortex can drive decisions

Millisecond-scale differences in neural activity in auditory cortex can drive decisions BRIEF COMMUNICATIONS 2008 Nature Publishing Group http://www.nature.com/natureneuroscience Millisecond-scale differences in neural activity in auditory cortex can drive decisions Yang Yang 1,2, Michael

More information

Reward prediction based on stimulus categorization in. primate lateral prefrontal cortex

Reward prediction based on stimulus categorization in. primate lateral prefrontal cortex Reward prediction based on stimulus categorization in primate lateral prefrontal cortex Xiaochuan Pan, Kosuke Sawa, Ichiro Tsuda, Minoro Tsukada, Masamichi Sakagami Supplementary Information This PDF file

More information

The Journal of Physiology Neuroscience

The Journal of Physiology Neuroscience J Physiol 595.4 (217) pp 1393 1412 1393 The Journal of Physiology Neuroscience Excitation of lateral habenula neurons as a neural mechanism underlying ethanol-induced conditioned taste aversion Shashank

More information

Lecturer: Rob van der Willigen 11/9/08

Lecturer: Rob van der Willigen 11/9/08 Auditory Perception - Detection versus Discrimination - Localization versus Discrimination - - Electrophysiological Measurements Psychophysical Measurements Three Approaches to Researching Audition physiology

More information

ANTECEDENT REINFORCEMENT CONTINGENCIES IN THE STIMULUS CONTROL OF AN A UDITORY DISCRIMINA TION' ROSEMARY PIERREL AND SCOT BLUE

ANTECEDENT REINFORCEMENT CONTINGENCIES IN THE STIMULUS CONTROL OF AN A UDITORY DISCRIMINA TION' ROSEMARY PIERREL AND SCOT BLUE JOURNAL OF THE EXPERIMENTAL ANALYSIS OF BEHAVIOR ANTECEDENT REINFORCEMENT CONTINGENCIES IN THE STIMULUS CONTROL OF AN A UDITORY DISCRIMINA TION' ROSEMARY PIERREL AND SCOT BLUE BROWN UNIVERSITY 1967, 10,

More information

KEY PECKING IN PIGEONS PRODUCED BY PAIRING KEYLIGHT WITH INACCESSIBLE GRAIN'

KEY PECKING IN PIGEONS PRODUCED BY PAIRING KEYLIGHT WITH INACCESSIBLE GRAIN' JOURNAL OF THE EXPERIMENTAL ANALYSIS OF BEHAVIOR 1975, 23, 199-206 NUMBER 2 (march) KEY PECKING IN PIGEONS PRODUCED BY PAIRING KEYLIGHT WITH INACCESSIBLE GRAIN' THOMAS R. ZENTALL AND DAVID E. HOGAN UNIVERSITY

More information

Lecturer: Rob van der Willigen 11/9/08

Lecturer: Rob van der Willigen 11/9/08 Auditory Perception - Detection versus Discrimination - Localization versus Discrimination - Electrophysiological Measurements - Psychophysical Measurements 1 Three Approaches to Researching Audition physiology

More information

Supplementary Information for Correlated input reveals coexisting coding schemes in a sensory cortex

Supplementary Information for Correlated input reveals coexisting coding schemes in a sensory cortex Supplementary Information for Correlated input reveals coexisting coding schemes in a sensory cortex Luc Estebanez 1,2 *, Sami El Boustani 1 *, Alain Destexhe 1, Daniel E. Shulz 1 1 Unité de Neurosciences,

More information

TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab

TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab This report details my summer research project for the REU Theoretical and Computational Neuroscience program as

More information

A Memory Model for Decision Processes in Pigeons

A Memory Model for Decision Processes in Pigeons From M. L. Commons, R.J. Herrnstein, & A.R. Wagner (Eds.). 1983. Quantitative Analyses of Behavior: Discrimination Processes. Cambridge, MA: Ballinger (Vol. IV, Chapter 1, pages 3-19). A Memory Model for

More information

Supplementary Information

Supplementary Information Supplementary Information The neural correlates of subjective value during intertemporal choice Joseph W. Kable and Paul W. Glimcher a 10 0 b 10 0 10 1 10 1 Discount rate k 10 2 Discount rate k 10 2 10

More information

Value transfer in a simultaneous discrimination by pigeons: The value of the S + is not specific to the simultaneous discrimination context

Value transfer in a simultaneous discrimination by pigeons: The value of the S + is not specific to the simultaneous discrimination context Animal Learning & Behavior 1998, 26 (3), 257 263 Value transfer in a simultaneous discrimination by pigeons: The value of the S + is not specific to the simultaneous discrimination context BRIGETTE R.

More information

HCS 7367 Speech Perception

HCS 7367 Speech Perception Long-term spectrum of speech HCS 7367 Speech Perception Connected speech Absolute threshold Males Dr. Peter Assmann Fall 212 Females Long-term spectrum of speech Vowels Males Females 2) Absolute threshold

More information

Facilitative Effects of the Ampakine CX516 on Short-Term Memory in Rats: Correlations with Hippocampal Neuronal Activity

Facilitative Effects of the Ampakine CX516 on Short-Term Memory in Rats: Correlations with Hippocampal Neuronal Activity The Journal of Neuroscience, April 1, 1998, 18(7):2748 2763 Facilitative Effects of the Ampakine CX516 on Short-Term Memory in Rats: Correlations with Hippocampal Neuronal Activity Robert E. Hampson, 1

More information

Supplemental Material

Supplemental Material Supplemental Material Recording technique Multi-unit activity (MUA) was recorded from electrodes that were chronically implanted (Teflon-coated platinum-iridium wires) in the primary visual cortex representing

More information

Unit 1 Exploring and Understanding Data

Unit 1 Exploring and Understanding Data Unit 1 Exploring and Understanding Data Area Principle Bar Chart Boxplot Conditional Distribution Dotplot Empirical Rule Five Number Summary Frequency Distribution Frequency Polygon Histogram Interquartile

More information

ABSTRACT 1. INTRODUCTION 2. ARTIFACT REJECTION ON RAW DATA

ABSTRACT 1. INTRODUCTION 2. ARTIFACT REJECTION ON RAW DATA AUTOMATIC ARTIFACT REJECTION FOR EEG DATA USING HIGH-ORDER STATISTICS AND INDEPENDENT COMPONENT ANALYSIS A. Delorme, S. Makeig, T. Sejnowski CNL, Salk Institute 11 N. Torrey Pines Road La Jolla, CA 917,

More information

Hippocampal Time Cells : Time versus Path Integration

Hippocampal Time Cells : Time versus Path Integration Article Hippocampal Time Cells : Time versus Path Integration Benjamin J. Kraus, 1, * Robert J. Robinson II, 1 John A. White, 2 Howard Eichenbaum, 1 and Michael E. Hasselmo 1 1 Center for Memory and Brain,

More information

Profound Context-Dependent Plasticity of Mitral Cell Responses in Olfactory Bulb

Profound Context-Dependent Plasticity of Mitral Cell Responses in Olfactory Bulb Profound Context-Dependent Plasticity of Mitral Cell Responses in Olfactory Bulb Wilder Doucette, Diego Restrepo * PLoS BIOLOGY Department of Cell and Developmental Biology, Neuroscience Program and Rocky

More information

TEMPORALLY SPECIFIC BLOCKING: TEST OF A COMPUTATIONAL MODEL. A Senior Honors Thesis Presented. Vanessa E. Castagna. June 1999

TEMPORALLY SPECIFIC BLOCKING: TEST OF A COMPUTATIONAL MODEL. A Senior Honors Thesis Presented. Vanessa E. Castagna. June 1999 TEMPORALLY SPECIFIC BLOCKING: TEST OF A COMPUTATIONAL MODEL A Senior Honors Thesis Presented By Vanessa E. Castagna June 999 999 by Vanessa E. Castagna ABSTRACT TEMPORALLY SPECIFIC BLOCKING: A TEST OF

More information

Spatiotemporal clustering of synchronized bursting events in neuronal networks

Spatiotemporal clustering of synchronized bursting events in neuronal networks Spatiotemporal clustering of synchronized bursting events in neuronal networks Uri Barkan a David Horn a,1 a School of Physics and Astronomy, Tel Aviv University, Tel Aviv 69978, Israel Abstract in vitro

More information

Neural correlates of multisensory cue integration in macaque MSTd

Neural correlates of multisensory cue integration in macaque MSTd Neural correlates of multisensory cue integration in macaque MSTd Yong Gu, Dora E Angelaki,3 & Gregory C DeAngelis 3 Human observers combine multiple sensory cues synergistically to achieve greater perceptual

More information

Feasibility Evaluation of a Novel Ultrasonic Method for Prosthetic Control ECE-492/3 Senior Design Project Fall 2011

Feasibility Evaluation of a Novel Ultrasonic Method for Prosthetic Control ECE-492/3 Senior Design Project Fall 2011 Feasibility Evaluation of a Novel Ultrasonic Method for Prosthetic Control ECE-492/3 Senior Design Project Fall 2011 Electrical and Computer Engineering Department Volgenau School of Engineering George

More information

A model of the interaction between mood and memory

A model of the interaction between mood and memory INSTITUTE OF PHYSICS PUBLISHING NETWORK: COMPUTATION IN NEURAL SYSTEMS Network: Comput. Neural Syst. 12 (2001) 89 109 www.iop.org/journals/ne PII: S0954-898X(01)22487-7 A model of the interaction between

More information

on both components of conc Fl Fl schedules, c and a were again less than 1.0. FI schedule when these were arranged concurrently.

on both components of conc Fl Fl schedules, c and a were again less than 1.0. FI schedule when these were arranged concurrently. JOURNAL OF THE EXPERIMENTAL ANALYSIS OF BEHAVIOR 1975, 24, 191-197 NUMBER 2 (SEPTEMBER) PERFORMANCE IN CONCURRENT INTERVAL SCHEDULES: A SYSTEMATIC REPLICATION' BRENDA LOBB AND M. C. DAVISON UNIVERSITY

More information

PCA Enhanced Kalman Filter for ECG Denoising

PCA Enhanced Kalman Filter for ECG Denoising IOSR Journal of Electronics & Communication Engineering (IOSR-JECE) ISSN(e) : 2278-1684 ISSN(p) : 2320-334X, PP 06-13 www.iosrjournals.org PCA Enhanced Kalman Filter for ECG Denoising Febina Ikbal 1, Prof.M.Mathurakani

More information

A Race Model of Perceptual Forced Choice Reaction Time

A Race Model of Perceptual Forced Choice Reaction Time A Race Model of Perceptual Forced Choice Reaction Time David E. Huber (dhuber@psyc.umd.edu) Department of Psychology, 1147 Biology/Psychology Building College Park, MD 2742 USA Denis Cousineau (Denis.Cousineau@UMontreal.CA)

More information

STUDIES OF WHEEL-RUNNING REINFORCEMENT: PARAMETERS OF HERRNSTEIN S (1970) RESPONSE-STRENGTH EQUATION VARY WITH SCHEDULE ORDER TERRY W.

STUDIES OF WHEEL-RUNNING REINFORCEMENT: PARAMETERS OF HERRNSTEIN S (1970) RESPONSE-STRENGTH EQUATION VARY WITH SCHEDULE ORDER TERRY W. JOURNAL OF THE EXPERIMENTAL ANALYSIS OF BEHAVIOR 2000, 73, 319 331 NUMBER 3(MAY) STUDIES OF WHEEL-RUNNING REINFORCEMENT: PARAMETERS OF HERRNSTEIN S (1970) RESPONSE-STRENGTH EQUATION VARY WITH SCHEDULE

More information

Mnemonic representations of transient stimuli and temporal sequences in the rodent hippocampus in vitro

Mnemonic representations of transient stimuli and temporal sequences in the rodent hippocampus in vitro Supplementary Material Mnemonic representations of transient stimuli and temporal sequences in the rodent hippocampus in vitro Robert. Hyde and en W. Strowbridge Mossy ell 1 Mossy ell Mossy ell 3 Stimulus

More information

Information-theoretic stimulus design for neurophysiology & psychophysics

Information-theoretic stimulus design for neurophysiology & psychophysics Information-theoretic stimulus design for neurophysiology & psychophysics Christopher DiMattina, PhD Assistant Professor of Psychology Florida Gulf Coast University 2 Optimal experimental design Part 1

More information

Mammogram Analysis: Tumor Classification

Mammogram Analysis: Tumor Classification Mammogram Analysis: Tumor Classification Term Project Report Geethapriya Raghavan geeragh@mail.utexas.edu EE 381K - Multidimensional Digital Signal Processing Spring 2005 Abstract Breast cancer is the

More information

7 Grip aperture and target shape

7 Grip aperture and target shape 7 Grip aperture and target shape Based on: Verheij R, Brenner E, Smeets JBJ. The influence of target object shape on maximum grip aperture in human grasping movements. Exp Brain Res, In revision 103 Introduction

More information

Comment by Delgutte and Anna. A. Dreyer (Eaton-Peabody Laboratory, Massachusetts Eye and Ear Infirmary, Boston, MA)

Comment by Delgutte and Anna. A. Dreyer (Eaton-Peabody Laboratory, Massachusetts Eye and Ear Infirmary, Boston, MA) Comments Comment by Delgutte and Anna. A. Dreyer (Eaton-Peabody Laboratory, Massachusetts Eye and Ear Infirmary, Boston, MA) Is phase locking to transposed stimuli as good as phase locking to low-frequency

More information

Representations of single and compound stimuli in negative and positive patterning

Representations of single and compound stimuli in negative and positive patterning Learning & Behavior 2009, 37 (3), 230-245 doi:10.3758/lb.37.3.230 Representations of single and compound stimuli in negative and positive patterning JUSTIN A. HARRIS, SABA A GHARA EI, AND CLINTON A. MOORE

More information

TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION

TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION Karin Zimmer, R. Duncan Luce and Wolfgang Ellermeier Institut für Kognitionsforschung der Universität Oldenburg, Germany Institute

More information

Supporting Online Material for

Supporting Online Material for www.sciencemag.org/cgi/content/full/319/5871/1849/dc1 Supporting Online Material for Rule Learning by Rats Robin A. Murphy,* Esther Mondragón, Victoria A. Murphy This PDF file includes: *To whom correspondence

More information

Sound Texture Classification Using Statistics from an Auditory Model

Sound Texture Classification Using Statistics from an Auditory Model Sound Texture Classification Using Statistics from an Auditory Model Gabriele Carotti-Sha Evan Penn Daniel Villamizar Electrical Engineering Email: gcarotti@stanford.edu Mangement Science & Engineering

More information

SUPPLEMENTAL MATERIAL

SUPPLEMENTAL MATERIAL 1 SUPPLEMENTAL MATERIAL Response time and signal detection time distributions SM Fig. 1. Correct response time (thick solid green curve) and error response time densities (dashed red curve), averaged across

More information

6. Unusual and Influential Data

6. Unusual and Influential Data Sociology 740 John ox Lecture Notes 6. Unusual and Influential Data Copyright 2014 by John ox Unusual and Influential Data 1 1. Introduction I Linear statistical models make strong assumptions about the

More information

The Role of Temporal Relationships in the Transfer of Conditioned Inhibition

The Role of Temporal Relationships in the Transfer of Conditioned Inhibition Denniston, James C., Robert P. Cole, and Ralph R. Miller. (1998). "The role of temporal relationships in the transfer of conditioned inhibition." Journal of Experimental Psychology: Animal Behavior Processes

More information

Morton-Style Factorial Coding of Color in Primary Visual Cortex

Morton-Style Factorial Coding of Color in Primary Visual Cortex Morton-Style Factorial Coding of Color in Primary Visual Cortex Javier R. Movellan Institute for Neural Computation University of California San Diego La Jolla, CA 92093-0515 movellan@inc.ucsd.edu Thomas

More information

Experience-dependent recovery of vision following chronic deprivation amblyopia. Hai-Yan He, Baisali Ray, Katie Dennis and Elizabeth M.

Experience-dependent recovery of vision following chronic deprivation amblyopia. Hai-Yan He, Baisali Ray, Katie Dennis and Elizabeth M. Experience-dependent recovery of vision following chronic deprivation amblyopia Hai-Yan He, Baisali Ray, Katie Dennis and Elizabeth M. Quinlan a 3. 2.5 2. 1.5.5 Deprived eye Non-deprived VCtx * * b 8.

More information

A Race Model of Perceptual Forced Choice Reaction Time

A Race Model of Perceptual Forced Choice Reaction Time A Race Model of Perceptual Forced Choice Reaction Time David E. Huber (dhuber@psych.colorado.edu) Department of Psychology, 1147 Biology/Psychology Building College Park, MD 2742 USA Denis Cousineau (Denis.Cousineau@UMontreal.CA)

More information

Nature Neuroscience: doi: /nn Supplementary Figure 1

Nature Neuroscience: doi: /nn Supplementary Figure 1 Supplementary Figure 1 Reward rate affects the decision to begin work. (a) Latency distributions are bimodal, and depend on reward rate. Very short latencies (early peak) preferentially occur when a greater

More information

CAROL 0. ECKERMAN UNIVERSITY OF NORTH CAROLINA. in which stimulus control developed was studied; of subjects differing in the probability value

CAROL 0. ECKERMAN UNIVERSITY OF NORTH CAROLINA. in which stimulus control developed was studied; of subjects differing in the probability value JOURNAL OF THE EXPERIMENTAL ANALYSIS OF BEHAVIOR 1969, 12, 551-559 NUMBER 4 (JULY) PROBABILITY OF REINFORCEMENT AND THE DEVELOPMENT OF STIMULUS CONTROL' CAROL 0. ECKERMAN UNIVERSITY OF NORTH CAROLINA Pigeons

More information

File name: Supplementary Information Description: Supplementary Figures, Supplementary Table and Supplementary References

File name: Supplementary Information Description: Supplementary Figures, Supplementary Table and Supplementary References File name: Supplementary Information Description: Supplementary Figures, Supplementary Table and Supplementary References File name: Supplementary Data 1 Description: Summary datasheets showing the spatial

More information